• Aucun résultat trouvé

Seismic Imaging of Thickened Lithosphere Resulting From Plume Pulsing Beneath Iceland

N/A
N/A
Protected

Academic year: 2021

Partager "Seismic Imaging of Thickened Lithosphere Resulting From Plume Pulsing Beneath Iceland"

Copied!
12
0
0

Texte intégral

(1)

HAL Id: insu-02919988

https://hal-insu.archives-ouvertes.fr/insu-02919988

Submitted on 24 Aug 2020

HAL is a multi-disciplinary open access

archive for the deposit and dissemination of

sci-entific research documents, whether they are

pub-lished or not. The documents may come from

teaching and research institutions in France or

abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est

destinée au dépôt et à la diffusion de documents

scientifiques de niveau recherche, publiés ou non,

émanant des établissements d’enseignement et de

recherche français ou étrangers, des laboratoires

publics ou privés.

Catherine Rychert, Nicholas Harmon, John Armitage

To cite this version:

Catherine Rychert, Nicholas Harmon, John Armitage. Seismic Imaging of Thickened Lithosphere

Resulting From Plume Pulsing Beneath Iceland. Geochemistry, Geophysics, Geosystems, AGU and

the Geochemical Society, 2018, 19 (6), pp.1789-1799. �10.1029/2018GC007501�. �insu-02919988�

(2)

RESEARCH ARTICLE

10.1029/2018GC007501

Seismic Imaging of Thickened Lithosphere Resulting From

Plume Pulsing Beneath Iceland

Catherine A. Rychert1 , Nicholas Harmon1 , and John J. Armitage2

1National Oceanography Centre Southampton, Ocean and Earth Sciences, University of Southampton, Southampton, UK, 2

Dynamique des Fluides Geologiques, Institut de Physique du Globe de Paris, Paris, France

Abstract

Ocean plates conductively cool and subside with seafloor age. Plate thickening with age is also predicted, and hot spots may cause thinning. However, both are debated and depend on the way the plate is defined. Determining the thickness of the plates along with the process that governs it has proven chal-lenging. We use S-to-P (Sp) receiver functions to image a strong, persistent LAB beneath Iceland where the mid-Atlantic Ridge interacts with a plume with hypothesized pulsating thermal anomaly. The plate is thick-est, up to 84 6 6 km, beneath lithosphere formed during times of hypothesized hotter plume temperatures and as thin as 61 6 6 km beneath regions formed during colder intervals. We performed geodynamic modeling to show that these plate thicknesses are inconsistent with a thermal lithosphere. Instead, periods of increased plume temperatures likely increased the melting depth, causing deeper depletion and dehy-dration, and creating a thicker plate. This suggests plate thickness is dictated by the conditions of plate formation.

1. Introduction

The lithosphere-asthenosphere boundary (LAB) represents the base of the tectonic plate, a fundamental boundary in plate tectonics. It represents the transition of the rigid plate to the weaker asthenosphere. Since it is well established that temperature plays a large role in the strength of materials, the LAB is typically assumed to follow an isotherm (Parsons & Sclater, 1977), with ocean plates progressively thickening as they cool and age (Parsons & Sclater, 1977) and thinning in the presence of hot spots (Detrick & Crough, 1978). In this model, seismic velocities are predicted to gradually decrease from the lithosphere to the astheno-sphere (Jackson & Faul, 2010; Rychert et al., 2010; Tharimena et al., 2017; Rychert & Harmon, 2018). However, several observations of sharp velocity gradients (Kawakatsu et al., 2009; Rychert et al., 2005; Rychert & Shearer, 2009) and discontinuities at constant depth beneath the oceans (Gaherty et al., 1999; Tan & Helm-berger, 2007; Tharimena et al., 2017) suggest a chemical boundary or a variation in melt might exist that could affect the strength of the mantle (Hirth & Kohlstedt, 1995; Jackson et al., 2006) and therefore influence the location of the LAB (Rychert et al., 2005). A compositional boundary might reflect the depth extent of depletion caused by melting during plate formation or subsequent melting due to hot spot anomalies (Gah-erty et al., 1999; Hall & Kincaid, 2003; Hirth & Kohlstedt, 1996; Jordan, 1978). Melt ponding, possibly con-trolled by the solidus (Sparks & Parmentier, 1991) or the melt-solid density contrast, might also determine the depth of the LAB (Sakamaki et al., 2013). However, the exact relationship between composition and/or melt variations and plate thickness has yet to be established.

Iceland is a unique opportunity to study the definition of the plate given its location above a thermal plume anomaly currently interacting with a mid-ocean ridge. Mid-ocean ridges are ideal locations to study the plate definition, given the simple and short history of the plate. Ocean islands like Iceland provide rare in situ constraints on the oceanic lithosphere, given the difficulties of ocean bottom seismic experiments. Hypothesized temporal variations in the magnitude of the thermal plume anomaly beneath Iceland may also be evaluated (Ito, 2001; Jones et al., 2002). Here we use Sp receiver functions to investigate discontinu-ity structure related to the LAB beneath Iceland. We also perform numerical modeling of mantle flow, tem-perature and melting beneath Iceland, and translate the thermal structure to predicted seismic velocity and receiver functions to test whether a purely thermally defined lithospheric plate may explain our observa-tions, or another mechanism is required.

Key Points:

S-to-P receiver functions image laterally variable lithospheric thickness beneath Iceland

Lithospheric thickness is greatest in regions formed during hypothesized periods of greater plume flux or temperature

Geodynamic and waveform modeling suggest melt beneath the plate, with thickness dictated by the conditions of formation

Correspondence to:

C. A. Rychert, c.rychert@soton.ac.uk

Citation:

Rychert, C. A., Harmon, N., & Armitage, J. J. (2018). Seismic imaging of thickened lithosphere resulting from plume pulsing beneath Iceland. Geochemistry, Geophysics, Geosystems, 19, 1789–1799. https://doi.org/10. 1029/2018GC007501

Received 20 FEB 2018 Accepted 4 MAY 2018

Accepted article online 12 MAY 2018 Published online 22 JUN 2018

VC2018. The Authors.

This is an open access article under the terms of the Creative Commons Attri-bution License, which permits use, dis-tribution and reproduction in any medium, provided the original work is properly cited.

(3)

2. Methods

2.1. Data Processing

We consider 4,265 waveforms recorded by stations in the IRIS database (www.iris.edu) from 1996 to 2005 and epicentral distances 55–808. We hand-pick the data, selecting the best 618 S-waves. We calculate receiver functions using two methods of Sp deconvolution, a multitaper (Helffrich, 2006; Rychert et al., 2012) method which is stable for individual waveforms and provides good 3-D structure and a simultaneous deconvolution (Rychert et al., 2007) which can verify the robustness of results. We experiment with a variety of frequency bands. A more limited band, typical for ocean islands (Li et al., 2004; Rychert et al., 2013), gives simpler waveforms, which we use for waveform modeling 0.05–0.14 Hz. The interpreted features of the stacked 3-D image are robust regardless of filtering and we present a wider band, 0.03–0.25 Hz, to demon-strate the highest resolution we could achieve. Deconvolved waveforms are multiplied by negative one, to match polarity of Ps receiver functions.

In the multitaper deconvolution, we eliminate individual unstable deconvolutions, characterized by ringing, where large amplitudes recur periodically for Sp delay times of 0 to 270 s (Rychert et al., 2007, 2012). This leaves 304 waveforms for the final multitaper imaging. The receiver functions are migrated to depth in 3-D. Migration model crustal thicknesses correspond to the value from the surface wave model (Li & Detrick, 2006) at the Sp conversion points at 30 km depth. The crustal shear velocity beneath the station from sur-face waves is assumed. Crustal Vp/Vs is assumed to be 1.75, between the median (1.74) and the mean (1.76) values for the island based on Ps and PsPs arrival times (Kumar et al., 2007), and within the range from sev-eral active source seismic studies (1.72–1.79) (Staples et al., 1997; Tryggvason & Bath, 1961; Weir et al., 2001). In the mantle, we assume a 1-D shear velocity structure from surface waves, and Vp 5 8.0 km/s. Multi-taper waveforms are binned at 1 km depth spacing on a 0.258 by 0.258 grid from 0 to 80 km depth and a 0.758 by 0.758 grid at deeper depths. The grid is smoothed with a radius corresponding to the Fresnel zone of the waveform. Only bins with greater than 3 waveforms where data amplitude exceeds the formal error bar (2 standard deviations from the mean) are included.

For the simultaneous deconvolution we divide the data into two large bins, one in the southwest (305 waves) and one in the northeast (313 waves). Large bins are necessary to attain a robust result with this method. Migration model parameters correspond to the 1-D average of the migration model described above, based on surface waves (Li & Detrick, 2006).

2.2. Synthetic Waveform Modeling

We forward model the multitaper and simultaneous deconvolution results assuming a 1-D model and using a propagator matrix method to generate synthetic seismograms (Keith & Crampin, 1977). The synthetics are processed and deconvolved as receiver functions using the same parameters as the data (Helffrich, 2006; Rychert et al., 2012). The migration model shear velocities are updated for consistency with the forward modeling. We assume the crust is a single layer when forward modeling, for simplicity. The dominant period of the waveforms is assumed to be 12 s and is determined by auto-convolution and considering the charac-ter of deconvolved waveforms.

Error bars on the simultaneous deconvolution waveforms represent 95% confidence as determined by a bootstrap with 100 repeats. Error bars on the multitaper deconvolutions represent the formal 95% confi-dence limits. Errors in the magnitude and sharpness of the velocity contrast are determined by changing the forward model to reach the limits of the bootstrap error bar. Error in depth to the discontinuities (65 km for the Moho and 66 km for the LAB) reflects variability caused by changing the migration model Vp/Vs assumption by 5%, leaving shear velocity fixed.

In filtered versions of the data (0.05–0.14 Hz), the Moho appears as a broad phase, likely owing to inter-ference between the Moho at 25 and 40 km and the phase from the base of the extrusives at 8–10 km. We do not model this complexity. Testing indicates that complex crustal structure does not greatly influ-ence our results from deeper phases. Changing the crustal shear velocity by 10% affects the estimated magnitude of the LAB phase by <1%. We also test the effect of assuming a different crustal thickness model (Kumar et al., 2007) for migration, but find it affected results from the LAB discontinuity by <2 km depth.

(4)

2.3. Geodynamic Modeling

We solve Stokes flow with the Boussinesq approximation (Armitage et al., 2008; Moresi et al., 1996; Nielsen & Hopper, 2004). We assume that the mantle deforms by dislocation creep (Levy & Jaupart, 2011), that the presence of partial melt leads to a weakening that has an exponential relationship between bulk viscosity and porosity (Kele-men et al., 1997), and that the dehydration of the solid matrix leads to a strengthening of the upper mantle (Hirth & Kohlstedt, 1996). This leads to the following rheological law:

g5vdehye2a/A1

nexp E1pV

nRT

 

_Ið12nÞ=n

where the various constants are given in Table 1. The parameter vdehyis equal to 1 until 2% melt is

gener-ated, where it increases to 10. This is to simulate the strengthening effect of the removal of volatile com-pounds on the solid mantle matrix. We choose to increase the strength by only one order of magnitude given the on-going debate for actual effect of water on the strength of mantle rocks (Fei et al., 2013). The Stokes solver applies a viscosity cut-off to avoid very large differences in viscosity within the numerical domain. The maximum viscosity is 1023Pa s, and the minimum is 1017Pa s.

We present a model of extension of the upper mantle, where the extension is driven by a divergent kinematic velocity of 10 mm/yr. The velocity is applied at 20 km depth, to approximate the effect of an increased crustal thickness on the lithosphere thickness. The velocity boundary conditions are of free slip and for temperature they are fixed at the top, 08C, and bottom, 1,4508C, and are of zero gradient at the sides. This is the model we present here, although we also tested models assuming potential tem-peratures from 1,4008C to1,6008C in 508C increments. To drive extension at 20 km depth requires a high model resolution, 2.7 3 0.7 km (1,024 3 1,024 elements), so that the velocity can be applied at the nodal points within the model domain. The center of extension is moved laterally and the resultant tem-perature and melt production is plotted after 6.2 My postmigration. This is to simulate the ridge jump at6–7 Ma.

For both models where the velocity is applied at the surface and at 20 km depth we find that the tempera-ture field is very similar, despite the change in the depth of the driving velocity condition. This is because the flow field within the upper 50 km is reasonably similar. The vertical profile of the model viscosity shows an increase in viscosity that is due to the reduction in temperature. Above a depth of 30 km, the predicted mantle flow is sufficiently slow that thermal diffusion dominates over advection, and the isotherms are con-trolled by heat conduction. In the models viscosity increases to 1023Pa s, which is the cut-off. For the model where the divergent velocity is applied on the surface this viscosity increase to the cut-off is gradual. How-ever, when the velocity is applied within the model domain, we create a region of reduced viscosity where the flow is imposed, as the rheology is a function of strain rate and viscosity reduces with increasing strain rates. Yet this difference in viscosity is too minor to alter the overall flow of the mantle and therefore signifi-cantly alter the thermal structure of the model lithosphere-asthenosphere system.

We convert the thermal structure to seismic velocities (Jackson & Faul, 2010), assuming a primarily olivine mantle. The method accounts for attenuation effects based on laboratory scalings and we assume a grain size of 20 mm. The seismic velocity structure is calculated for each element in the model. We calculate the predicted receiver functions from the seismic velocities using a 1-D reflectivity code (Shearer & Orcutt, 1987) assuming a 20 km thick crust in averaging the velocity structure over 50 km wide bins. We use the same frequency band used in our study 0.03–0.25 Hz.

3. Results

Sp receiver functions image a shallow positive phase at 3–9 km depth (Figure 1). We also image a deeper positive phase that likely represents the Moho at 25–40 6 5 km depth (Figures 1 and 2). The two phases are pervasive across the region except at the intersection of the rift segments in the center of the island where we image a single positive discontinuity at 14 6 5 km depth (area of black shading, Figure 2).

Table 1

Geodynamic Model Parameters

Parameter Value Units Reference

a 45 Kelemen et al. (1997)

A 8.64 3 10212

E 523 kJ mol21 Korenaga and Karato (2008)

n 3.6 Korenaga and Karato (2008)

(5)

We image a negative phase, velocity decrease with depth at 61– 84 6 6 km depth in the multitaper deconvolution (Figures 1 and 3). Beneath the ridge axis the phase is shallow, 60–70 km (Figure 3). The phase is imaged deeper just outside the ridge centered Quater-nary volcanic area, with the deepest realizations approximately cen-tered on the 9 Ma isochron (Figure 3). The receiver function1-D profile in this region has two negative phases (Figure 1b), although the deeper phase is the most robust, interpretable feature, getting more pronounced when filtered to longer period. The phase is then shallower, 60–68 km, beneath oldest aged lithosphere in the north-west. The depth variations are independent of those of the Moho (Figures 2 and 3). The phase represents a strong velocity decrease with depth, 15 6 7% beneath a 9 My old plate and 10 6 5% beneath the South Iceland Volcanic Zone (SIVZ), as suggested by synthetic waveform modeling (Figure 4). The phase may be associated with the lithosphere-asthenosphere boundary, which will be assessed in subsequent sections.

A positive phase, velocity increase in depth is imaged sporadically in the multitaper cross sections at 80–90 km depth. The phase appears in the locations of the slowest seismic velocity anomalies from

Figure 1. Map of the study region and cross sections through the migrated Sp multitaper receiver function model. (a) Color contours show lithospheric age. Iceland coast outlined in black. Some of the active volcanic zones are labeled with abbreviations as follows: SNVZ, Snaefellsnes Volcanic Zone; MIVZ, Mid-Iceland Volcanic Zone; SIVZ, South Iceland Volcanic Zone; RRZ, Reykjanes Rift Zone; NIRZ, North Iceland Rift Zone. Extinct volcanic zones: SSRZ, Snæfellsnes-Skagi Rift Zone. Holocene volcanos are indicated with black triangles (Venzke, 2013). (b, c) Cross sections through the migrated Sp multi-taper receiver function model. Colors indicate the polarity of seismic discontinuities from receiver functions, positive, velocity increase with depth (red) and negative, velocity decrease with depth (blue). Open circles at 150 km depth are spaced at 100 km and plotted as reference points. Topography is plotted at top, with exaggeration. Regions with insuffi-cient data are shaded grey. Inverted red triangles show seismic station locations. Black contours show shear wave velocity from surface waves (Li & Burke, 2006), as labeled. Green line shows the 1,0008C isotherm from the geodynamic model. Horizontal dashed black lines indicate 20% depletion for a range of mantle potential temperatures, as labeled. Error in

depth of the LAB phase is6 km, as discussed in section 3.

Figure 2. Depth to the Moho from the multitaper method, map view. Bins with fewer than 3 waveforms are shaded grey. Black lines outline Iceland coast. Dark grey lines show age contours: 0, 2, 7, 8, and 12 My, as labeled. Cross section lines of Figure 1 are drawn as a thick black horizontal and vertical lines. Red and green inverted triangles show waveform modeling locations (Figure 4).

Error in depth of the Moho is5 km, as discussed in section 3. Black shading

shows the area where the Moho phase at similar depths (25–40 km) is not detected. Depths are determined automatically from migrated receiver function model.

(6)

surface waves (Figure 1). It is typically imaged at the deeper limit of the 3.9 km/s contour where negative LAB phase is also shallow (Figure 1c).

A positive phase, velocity increase with depth is also imaged at 150 km in both the large northeastern simultaneous deconvolution bin (Figure 5) and east of the island in the 3-D multitaper model (Fig-ure 1). Modeling the simultaneous deconvolution suggests a 4 6 1% velocity increase (Figure 5), although the data included averages over a large lateral area and therefore may underpredict the magnitude. Modeling the phase from the multitaper deconvolution is challenging given that it is at the edge of the model, and structure directly above it is not well-resolved (Figures 1 and 6). We model the phase assuming a shallow discontinuity structure from the area adjacent to the discon-tinuity and find a large17%, velocity increase. Therefore, the discon-tinuity likely represents a 4–17% velocity increase with depth. The forward modeling suggests that the velocity discontinuities could be sharp or occur over up to20 km depth without changing the shape of the synthetic receiver function or as wide as 50 km depth before reaching the bounds of the error bars from bootstrap.

Numerical modeling of mantle flow and temperature at the ridge pre-dicts progressive thickening of the thermally defined plate with distance from the ridge (Figure 7). Translat-ing these temperatures to seismic velocity results in increasTranslat-ing velocities with distance from the ridge, as well as an increasing thickness to the seismically fast lid. Predicted receiver functions include a Moho. How-ever, no discernible LAB phase is predicted owing to the gradual nature of the velocity gradient at the base of the thermally defined plate. A negative feature, follows just beneath the Moho, as a sidelobe artifact. The increase in Moho amplitude away from the ridge axis occurs because the lithosphere is progressively faster.

Figure 3. Depth to the negative discontinuity (LAB) from the multitaper method, map view. Bins with fewer than 3 waveforms or amplitudes less than 0.05 are shaded grey. Black lines outline Iceland coast. Dark grey lines show age contours: 0, 2, 7, 8, and 12 My as labeled. Cross section lines of Figure 1 are drawn as thick black horizontal and vertical lines. Red and green inverted trian-gles show waveform modeling locations (Figure 4). Error in depth of the LAB

phase is6 km, as discussed in section 3. Depths are determined automatically

from migrated receiver function model.

Figure 4. Waveform modeling of Sp 3-D multitaper model (black) beneath the ridge and beneath the 9 My age contour (locations of triangles in Figure 3) are compared to synthetics from forward modeling (red and green lines). Correspond-ing shear velocity profiles are shown to the left of the data/synthetics plots. Error bars indicatCorrespond-ing 95% confidence from bootstrap are shown in grey.

(7)

4. Discussion

The shallow positive phase at 3–9 km agrees well with the strong shallow velocity gradients from 0 to 8– 10 km in ambient noise tomography and interpreted as the transition from fractured extrusive igneous rocks typically associated with layer two of the oceanic Moho to the deeper intrusive crust (Green et al., 2017). The deeper positive phase at 25–40 km depth likely represents the Moho. In the southern part of the North Iceland Rift Zone (NIRZ) where only a single crustal phase is imaged at intermediate depths,14 km, crustal structure could be more simple (single layer), or the Moho phase may be shallow, thus interfering with the phase from the base of the extrusives. Alternately the deeper Moho phase is weak or not existent and the extrusive layer is very thick.

Our Moho depths are in general agreement with the range of values reported from Ps (Ps) imaging and sur-face waves,15–42 km (Darbyshire et al., 2000; Kumar et al., 2007; Li & Detrick, 2006), even if details of the

Figure 5. Waveform modeling of Sp data from the simultaneous deconvolution. Sp data (black) binned by conversion point bins at 100 km depth into 2 bins (red and green dots in inset map) are compared to synthetics from forward model-ing for the southwest (green) and northeast (red). Correspondmodel-ing shear velocity profiles are shown to the left of the data/ synthetics plots. Error bars indicating 95% confidence from bootstrap are shown in grey.

Figure 6. Hit count map for the multitaper method at 75 (left), 100 (middle), and 150 (right) km depth. The number of waveforms averaged in each bin is shown by grey scale, 0 (white) to up to 80 (black). Coastlines are outlined in black. Black line indicates cross section location.

(8)

models are all slightly different. The general trend of increasing crustal thickness away from the ridge is in agreement with a similar trend in Ps receiver functions from the northern ridge segment (Darbyshire et al., 2000). Beneath the large B€ardarbunga volcano (longi-tude 5 217.58, latti(longi-tude 5 64.68) our 35 km thick crust agrees with the 32 km thick crust from Ps receiver functions (Kumar et al., 2007). It is also within error of the 40 km thickness from refraction (Darbyshire et al., 1998), and Ps receiver function modeling that suggested a Moho velocity gradient from 30 to 40 km (Darbyshire et al., 2000). Although there is general agreement, variability at individual locations could be owing to complications of Ps from reverberated phases related to the shallower discontinuity at 3–9 km depth and/or lower lateral resolution in Sp in comparison to Ps.

The depth of the negative phase agrees with a previous Sp receiver function study which found a decrease in velocity with depth at 80 km depth beneath Iceland (Kumar et al., 2005). It is also in agree-ment with the location of the gradual drop in velocity from 50 to 100 km depth in surface wave velocity across the region and a fast lid that persists throughout (Li & Detrick, 2006). Our result provides higher lateral resolution (25–50 km) than these previous studies (100– 250 km). The modeled velocity contrasts from the multitaper (15 6 7% and 10 6 5% drops) are in agreement and in some cases slightly greater than the magnitude of the surface wave drops (9% beneath the island and5% near the ridge axis). Larger receiver func-tion results are likely explained by lateral or vertical smoothing in the surface waves.

The imaged negative phase is not consistent with predictions for a purely thermal model. Although a negative velocity gradient is pre-dicted beneath the seismically fast lithospheric lid (Figure 7b), the velocity gradient is too gradual, and no strong receiver function phase is predicted (Figure 7c). Particularly beneath the ridge, the seismically fast lithospheric lid is nearly nonexistent and the Moho is weak (Figure 7b). In addition, the depth of the observed negative phase does not correspond to the depth variations of the thermal model. It is deeper than the center of the negative velocity gradient predicted for a ther-mal model by 20–35 km beneath most the island (Figures 1 and 7b). Its depth variability is similarly more complicated than isotherm pat-terns for a conductively cooling lithosphere, which predict simple thickening with age (blue phase versus green line, Figure 1). A thermal plume could perturb progressive aging, thinning the lithosphere. However, our areas of thinned lithosphere do not correspond to either the expected plume location (Li & Detrick, 2006) or locations formed during hypothesized periods of hotter plume temperature (Jones et al., 2014). Another mechanism is required to explain the magnitude, sharpness, absolute depth, and lateral depth variations of the imaged velocity drop.

An additional mechanism, such as hydration, melting, or anisotropy can further decrease seismic velocities, in comparison to a thermally defined model. Bulk composition variations cannot explain the magni-tude of the seismic velocity drop alone (<2%) (Schutt & Lesher, 2006). An increase in hydration with depth could produce a large velocity reduction if hypothesized grain boundary sliding mechanisms are activated (Karato et al., 2015). Although, a hydrated mantle beneath Iceland is not likely since hydration would parti-tion into the high degrees of mantle melting that is thought to occur beneath Iceland. In addiparti-tion, recent

Figure 7. Geodynamic model of extension for the upper mantle. (a) Temperature, (b) seismic shear wave velocity, and (c) receiver function predictions are shown. The model shows results 6.2 Myr after a 100 km lateral shift in extension for the case where the divergent velocity condition is applied at 20 km depth. This model has a resolution of 1,024 by 1,024 elements (2.7 km laterally by 0.7 km in depth).

(9)

experimental work suggests that hydration does not impact seismic velocity (Cline et al., 2018). A change in radial anisotropy with depth can only explain a small (< 2%) apparent Sp velocity contrast, far smaller than our observations (Rychert & Harmon, 2017). Alternatively, a small amount of partial melting beneath the dis-continuity could easily explain the observed magnitude of the contrast (Clark & Lesher, 2017; Hammond & Humphreys, 2000), and it is suggested that melt naturally ponds in this depth range (Sakamaki et al., 2013). In this case, the observed phase would necessarily represent the LAB, since asthenospheric melt at concen-trations imageable by seismic waves (1%) (Hammond & Humphreys, 2000) or more (Clark & Lesher, 2017) is expected to significantly weaken the mantle (Hirth & Kohlstedt, 1995; Jackson et al., 2006).

However, to also achieve the observed topography on the discontinuity may require a combination of mechanisms. One possibility is that a compositional boundary dictates the depth at which melt ponds in the mantle. If the overlying lithosphere is compositionally depleted and dehydrated, for instance in clino-pyroxene, which occurs at 20% depletion, it could strengthen the mantle and create a permeability boundary below which melt ponds (Sparks & Parmentier, 1991). Steady state solid mantle flow and melting would not occur shallower than the discontinuity owing to the strength of the depleted lithosphere and lack of fusible material, which similarly prevents the ascent of melt (Ito et al., 1999). In numerical models, the viscous depleted layer also prevents shallow melting and can produce the relatively modest crustal thickness observed in Iceland (Ito et al., 1999). Steady state numerical models of Iceland predict a viscous layer with a base that varies from 78 to 65 km depth in the spreading direction, in general agreement with our range of observed LAB depths (Ito et al., 1999). The depth variability is smaller than our observations and also with an opposite sense, thicker beneath the ridge. Nonsteady state behavior, like the model pro-posed here with a pulsing plume could result in greater topography, and provide a better match with our observations. For instance, this boundary would occur at 60 km depth assuming a mantle potential temper-ature of1,5158C. It would occur deeper during times of increased plume thermal anomalies, for instance 84 km for 1,6008C. The depth variability would explain the observed undulations in our negative polarity phase.

The observation of a persistent thick lithosphere beneath the ridge is very different from constraints from nonhot spot affected ridges. For instance, at the fast-spreading East Pacific Rise surface waves detect no seismically fast lithospheric lid (Harmon et al., 2009). At the intermediate spreading Juan de Fuca and Gorda Ridges receiver functions and surface waves similarly image seismically fast lithosphere that thins toward the ridge axis, reaching <25 km thickness (Audet, 2016; Bell et al., 2016; Rychert et al., 2018). Low seismic velocity and resistivity in these cases are typically interpreted as zones of shallow partial melt (Bell et al., 2016; Harmon et al., 2009; Key et al., 2013; Rychert et al., 2018; Tian et al., 2013).

The observation of a persistent thick lithosphere is similar to previous imaging from hot spot affected ridges. For example, beneath the intermediate-spreading Nazca Spreading Centre just north of the Galapa-gos hot spot a deep velocity drop at 40–70 km is imaged with surface waves (Villagomez et al., 2007) and receiver functions (Byrnes et al., 2015; Rychert et al., 2014) across the ridge. These hot spot affected results are typically interpreted as depleted lithospheric material with additional strength (Byrnes et al., 2015; Rychert et al., 2014; Villagomez et al., 2007), lending further support of a compositional lithosphere. Variations in the thickness of a compositionally depleted layer could be caused by different spreading and upwelling rates or different potential temperatures at the time of formation. Temporal variations in spread-ing rate are not evidenced here. More likely, episodes of increased thermal plume anomalies explain the observed lithospheric thickness variations. Periods of increased plume temperatures or flux could cause increased melting depths that are reflected in thicker regions of depletion and dehydration, and also thicker tectonic plates (Ito, 2001; Yamamoto & Morgan, 2009). Alternatively, propagation of small scale convective instabilities could periodically increase upwelling (Martinez & Hey, 2017). Indeed, our results show a deep negative phase beneath lithosphere formed during time period of hypothesized high plume temperatures and or flux, 9 Ma (Jones et al., 2002) (Figure 3), with thinner lithosphere in other regions. This suggests that tectonic plate thickness is likely dictated by the conditions of plate formation.

We illustrate a magmatic-tectonic scenario that could explain the observed variation in lithospheric thick-ness (Figures 1 and 3) in Figure 8. We assume plume pulsing beneath the ridge at 9 and 4 Ma based on modeling of crustal thickness and geochemical variations at the nearby Reykjanes Ridge (Jones et al., 2014) and use magnetic isochrones and associated spreading rates (Ivarsson, 1992; Martin et al., 2011) to track the

(10)

formation and evolution of the thickened lithosphere in the region. At 9–8 Ma, a plume pulse created a 100 km wide region of thickened lithosphere across the active ridge (Figure 8). The ridge proceeded to spread at 10 mm/yr. At 7–6 Ma, a new ridge formed to the east and also rifted the pre-existing thickened lithosphere, spreading at 20 mm/yr. The western ridge system died off. At 5–4 Ma, a new plume pulse thick-ened a100 km wide region beneath the active ridge, creating a nearly continuously thickened lithosphere beneath 4–9 My old lithosphere. From 3 to 0 Ma, the thickened lithosphere rifted apart at 20 mm/yr, thin-ning the area beneath the ridge. The resulting predicted depth to the LAB from the tectonic model (Figure 8) is strikingly similar to that observed in our result (Figure 3).

Additional positive phases at depth may represent velocity increases related to the reduction of melt concen-trations with depth. The positive phase at 80–90 km may represent the base of a highly concentrated melt layer, owing to ponding (Sakamaki et al., 2013). The deeper positive phase near 150 6 10 km depth agrees with a similar boundary imaged by a previous Sp receiver function result at 135 6 5 km (Vinnik et al., 2005) and at similar depths beneath other plumes globally (Li et al., 2000; Rychert et al., 2013, 2014). This boundary is likely related to the location of the thermal plume anomaly. This location is just east of the central volcanic region. Possible explanations include mantle plume ascent that is not perfectly vertical (Ballmer et al., 2015;

Figure 8. Schematic of the evolution of lithospheric thickness beneath Iceland. The left column shows map views, 8 Ma to present at 2 Ma intervals (top to bottom) with isochons (Ivarsson, 1992). Isochrons are as labeled. Blue areas show hypothe-sized regions of thickened lithosphere caused by increased plume anomaly at 9 (light blue) and 4 Ma (dark blue). Cross sec-tion indicated as thick black line on 0 Ma map. Right column shows cross secsec-tions with lithosphere in blue. Triangles show active (red) and extinct (black) spreading ridge. Red hatched regions indicate areas of hypothesized melt generation.

(11)

Steinberger & Antretter, 2006; Yamamoto & Morgan, 2009) and/or mantle plume deflection at depths shal-lower than 150 km (Yamamoto & Morgan, 2009). Indeed, a plume that is located to the east of Iceland at man-tle depths is compatible with progressive eastward ridge jumps occurring over the past 14 My (Ivarsson, 1992; Martin et al., 2011). Although the hypothesized location of the thermal plume anomaly is likely outside the well-resolved area of surface wave inversions, the strongest velocity anomalies in the 75–150 km depth range occur just north or east of Iceland (Li & Detrick, 2006), in agreement with our result.

References

Armitage, J. J., Henstock, T. J., Minshull, T. A., & Hopper, J. R. (2008). Modelling the composition of melts formed during continental breakup of the Southeast Greenland margin. Earth and Planetary Science Letters, 269(1–2), 248–258.

Audet, P. (2016). Receiver functions using OBS data: Promises and limitations from numerical modelling and examples from the Cascadia Initiative. Geophysical Journal International, 205(3), 1740–1755.

Ballmer, M. D., Ito, G., & Cheng, C. (2015). Asymmetric dynamical behavior of thermochemical plumes and implications for Hawaiian Lava composition. Geophysical Monograph Series, 208, 35–57.

Bell, S., Ruan, Y. Y., & Forsyth, D. W. (2016). Ridge asymmetry and deep aqueous alteration at the trench observed from Rayleigh wave tomography of the Juan de Fuca plate. Journal of Geophysical Research, 121, 7298–7321. https://doi.org/10.1002/2016JB012990 Byrnes, J. S., Hooft, E. E. E., Toomey, D. R., Villagomez, D. R., Geist, D. J., & Solomon, S. C. (2015). An upper mantle seismic discontinuity

beneath the Galapagos Archipelago and its implications for studies of the lithosphere-asthenosphere boundary. Geochemistry, Geophys-ics, Geosystems, 16, 1070–1088. https://doi.org/10.1002/2014GC005694

Clark, A. N., & Lesher, C. E. (2017). Elastic properties of silicate melts: Implications for low velocity zones at the lithosphere-asthenosphere boundary. Science Advances, 3(12), e1701312.

Cline, C. J., Faul, U. H., David, E. C., Berry, A. J., & Jackson, I. (2018). Redox-influenced seismic properties of uppermantle olivine. Nature, 555(7696), 355.

Darbyshire, F. A., Bjarnason, I. T., White, R. S., & Flovenz, O. G. (1998). Crustal structure above the Iceland mantle plume imaged by the ICE-MELT refraction profile. Geophysical Journal International, 135(3), 1131–1149.

Darbyshire, F. A., Priestley, K. F., White, R. S., Stefansson, R., Gudmundsson, G. B., & Jakobsdottir, S. S. (2000). Crustal structure of central and northern Iceland from analysis of teleseismic receiver functions. Geophysical Journal International, 143(1), 163–184.

Detrick, R., & Crough, S. (1978). Island subsidence, hot spots, and lithospheric thinning. Journal of Geophysical Research, 83(B3), 1236–1244. Fei, H. Z., Wiedenbeck, M., Yamazaki, D., & Katsura, T. (2013). Small effect of water on upper-mantle rheology based on silicon self-diffusion

coefficients. Nature, 498(7453), 213.

Gaherty, J. B., Kato, M., & Jordan, T. H. (1999). Seismological structure of the upper mantle: A regional comparison of seismic layering. Phys-ics of the Earth and Planetary Interiors, 110(1–2), 21–41.

Green, R. G., Priestley, K. F., & White, R. S. (2017). Ambient noise tomography reveals upper crustal structure of Icelandic rifts. Earth and Planetary Science Letters, 466, 20–31.

Hall, P. S., & Kincaid, C. (2003). Melting, dehydration, and the dynamics of off-axis plume-ridge interaction. Geochemistry, Geophysics, Geo-systems, 4(9), 8510. https://doi.org/10.1029/2003GC000567

Hammond, W. C., & Humphreys, E. D. (2000). Upper mantle seismic wave velocity: Effects of realistic partial melt geometries. Journal of Geo-physical Research, 105(B5), 10975–10986.

Harmon, N., Forsyth, D. W., & Weeraratne, D. S. (2009). Thickening of young Pacific lithosphere from high-resolution Rayleigh wave tomog-raphy: A test of the conductive cooling model. Earth and Planetary Science Letters, 278(1–2), 96–106.

Helffrich, G. (2006). Extended-time multitaper frequency domain cross-correlation receiver-function estimation. Bulletin of the Seismological Society of America, 96(1), 344–347.

Hirth, G., & Kohlstedt, D. L. (1995). Experimental constraints on the dynamics of the partially molten upper-mantle. 2. Deformation in the dislocation creep regime. Journal of Geophysical Research, 100(B8), 15441–15449.

Hirth, G., & Kohlstedt, D. L. (1996). Water in the oceanic upper mantle: Implications for rheology, melt extraction and the evolution of the lithosphere. Earth and Planetary Science Letters, 144(1–2), 93–108.

Ito, G. (2001). Reykjanes ‘V’-shaped ridges originating from a pulsing and dehydrating mantle plume. Nature, 411(6838), 681–684. Ito, G., Shen, Y., Hirth, G., & Wolfe, C. J. (1999). Mantle flow, melting, and dehydration of the Iceland mantle plume. Earth and Planetary

Sci-ence Letters, 165(1), 81–96.

Ivarsson, G. (1992). Geology and petrochemistry of the Torfajokull central volcano in Central South Iceland, in association with the Icelandic Hot Spot and Rift Zones (PhD dissertation, 332 p.). Honolulu, HI: University of Hawaii.

Jackson, I., & Faul, U. H. (2010). Grainsize-sensitive viscoelastic relaxation in olivine: Towards a robust laboratory-based model for seismo-logical application. Physics of the Earth and Planetary Interiors, 183(1–2), 151–163.

Jackson, I., Faul, U. H., Fitz Gerald, J. D., & Morris, S. J. S. (2006). Contrasting viscoelastic behavior of melt-free and melt-bearing olivine: Implications for the nature of grain-boundary sliding. Materials Science and Engineering A, 442(1–2), 170–174.

Jones, S. M., Murton, B. J., Fitton, J. G., White, N. J., Maclennan, J., & Walters, R. L. (2014). A joint geochemical-geophysical record of time-dependent mantle convection south of Iceland. Earth and Planetary Science Letters, 386, 86–97.

Jones, S. M., White, N., & Maclennan, J. (2002). V-shaped ridges around Iceland: Implications for spatial and temporal patterns of mantle convection. Geochemistry, Geophysics, Geosystems, 3(10), 1059. https://doi.org/10.1029/2002GC000361

Jordan, T. H. (1978). Composition and development of continental tectosphere. Nature, 274(5671), 544–548.

Karato, S. I., Olugboji, T., & Park, J. (2015). Mechanisms and geologic significance of the mid-lithosphere discontinuity in the continents. Nature Geoscience, 8(7), 509–514.

Kawakatsu, H., Kumar, P., Takei, Y., Shinohara, M., Kanazawa, T., Araki, E., et al. (2009). Seismic evidence for sharp lithosphere-asthenosphere boundaries of oceanic plates. Science, 324(5926), 499–502.

Keith, C. M., & Crampin, S. (1977). Seismic body waves in anisotropic media—Synthetic seismograms. Geophysical Journal of the Royal Astro-nomical Society, 49(1), 225–243.

Kelemen, P. B., Hirth, G., Shimizu, N., Spiegelman, M., & Dick, H. J. B. (1997). A review of melt migration processes in the adiabatically upwelling mantle beneath oceanic spreading ridges. Philosophical Transactions of the Royal Society A, 355(1723), 283–318.

Acknowledgments

We thank A. Li for providing her surface wave model and Reidar Tronnes for providing a local age contour map. N. Harmon and C. Rychert acknowledge funding from the Natural Environment Research Council (NE/K000985/1 and NE/ M003507/1) and the European Research Council (GA 638665). J. Armitage is funded through the French Agence National de la Recherche, Accueil de Chercheurs de Haut Niveau call, grant acronym ‘‘InterRift’’.oˆ The seismic data used in this study are available online at the Incorporated Research Institutions for Seismology (IRIS) data management center (http://ds.iris.edu/ds/nodes/ dmc). The authors declare that they have no competing financial interests.

(12)

Key, K., Constable, S., Liu, L., & Pommier, A. (2013). Electrical image of passive mantel upwelling beneath the northern east Pacific Rise. Nature, 495(7442), 499–502.

Korenaga, J., & Karato, S. I. (2008). A new analysis of experimental data on olivine rheology. Journal of Geophysical Research, 113, B02403. https://doi.org/10.1029/2007JB005100

Kumar, P., Kind, R., Hanka, W., Wylegalla, K., Reigber, C., Yuan, X., et al. (2005). The lithosphere-asthenosphere boundary in the North-West Atlantic region. Earth and Planetary Science Letters, 236(1–2), 249–257.

Kumar, P., Kind, R., Priestley, K., & Dahl-Jensen, T. (2007). Crustal structure of Iceland and Greenland from receiver function studies. Journal of Geophysical Research, 112, B03301. https://doi.org/10.1029/2005JB003991

Levy, F., & Jaupart, C. (2011). Temperature and rheological properties of the mantle beneath the North American craton from an analysis of heat flux and seismic data. Journal of Geophysical Research, 116, B01408. https://doi.org/10.1029/2010JB007726

Li, A. B., & Burke, K. (2006). Upper mantle structure of southern Africa from Rayleigh wave tomography. Journal of Geophysical Research, 111, B10303. https://doi.org/10310.11029/12006JB004321

Li, A. B., & Detrick, R. S. (2006). Seismic structure of Iceland from Rayleigh wave inversions and geodynamic implications. Earth and Plane-tary Science Letters, 241(3–4), 901–912.

Li, X., Kind, R., Priestley, K., Sobolev, S. V., Tilmann, F., Yuan, X., et al. (2000). Mapping the Hawaiian plume conduit with converted seismic waves. Nature, 405(6789), 938–941.

Li, X. Q., Kind, R., Yuan, X. H., Wolbern, I., & Hanka, W. (2004). Rejuvenation of the lithosphere by the Hawaiian plume. Nature, 427(6977), 827–829. Martin, E., Paquette, J. L., Bosse, V., Ruffet, G., Tiepolo, M., & Sigmarsson, O. (2011). Geodynamics of rift-plume interaction in Iceland as

con-strained by new Ar-40/Ar-39 and in situ U-Pb zircon ages. Earth and Planetary Science Letters, 311(1–2), 28–38.

Martinez, F., & Hey, R. (2017). Propagating buoyant mantle upwelling on the Reykjanes Ridge. Earth and Planetary Science Letters, 457, 10–22. Moresi, L., Zhong, S. J., & Gurnis, M. (1996). The accuracy of finite element solutions of Stokes’ flow with strongly varying viscosity. Physics

of the Earth and Planetary Interiors, 97(1–4), 83–94.

Nielsen, T. K., & Hopper, J. R. (2004). From rift to drift: Mantle melting during continental breakup. Geochemistry, Geophysics, Geosystems, 5, Q07003. https://doi.org/10.1029/2003GC000662

Parsons, B., & Sclater, J. G. (1977). Analysis of variation of ocean-floor bathymetry and heat-flow with age. Journal of Geophysical Research, 82(5), 803–827.

Rychert, C. A., Fischer, K. M., & Rondenay, S. (2005). A sharp lithosphere-asthenosphere boundary imaged beneath eastern North America. Nature, 436(7050), 542–545.

Rychert, C. A., & Harmon, N. (2017). Constraints on the anisotropic contributions to velocity discontinuities at60 km depth beneath the Pacific. Geochemistry, Geophysics, Geosystems, 18, 2855. https://doi.org/10.1002/2017GC006850

Rychert, C. A., Harmon, N., & Ebinger, C. (2014). Receiver function imaging of lithospheric structure and the onset of melting beneath the Galapagos Archipelago. Earth and Planetary Science Letters, 388, 156–165.

Rychert, C. A., Harmon, N., & Tharimena, S. (2018). Scattered wave imaging of the oceanic plate in Cascadia. Science Advances, 4(2), eaao1908.

Rychert, C. A., Hammond, J. O. S., Harmon, N., Kendall, J. M., Keir, D., Ebinger, C., et al. (2012). Volcanism in the Afar Rift sustained by decom-pression melting with minimal plume influence. Nature Geoscience, 5(6), 406–409.

Rychert, C. A., Laske, G., Harmon, N., & Shearer, P. M. (2013). Seismic imaging of melting in a displaced Hawaiian Plume. Nature Geoscience, 6(8), 657–660.

Rychert, C. A., Rondenay, S., & Fischer, K. M. (2007). P-to-S and S-to-P imaging of a sharp lithosphere-asthenosphere boundary beneath eastern North America. Journal of Geophysical Research, 112, B08314. https://doi.org/10.1029/2006JB004619

Rychert, C. A., & Shearer, P. M. (2009). A global view of the lithosphere-asthenosphere boundary. Science, 324(5926), 495–498.

Rychert, C. A., & Harmon, N. (2018). Predictions and observations for the oceanic lithosphere from S-to-P receiver functions and SS precur-sors. Geophysical Research Letters, 45. https://doi.org/10.1029/2018GL077675

Rychert, C. A., Shearer, P. M., & Fischer, K. M. (2010). Scattered wave imaging of the lithosphere-asthenosphere boundary. Lithos, 120(1–2), 173–185. Sakamaki, T., Suzuki, A., Ohtani, E., Terasaki, H., Urakawa, S., Katayama, Y., et al. (2013). Ponded melt at the boundary between the

litho-sphere and asthenolitho-sphere. Nature Geoscience, 6(12), 1041–1044.

Schutt, D. L., & Lesher, C. E. (2006). Effects of melt depletion on the density and seismic velocity of garnet and spinel lherzolite. Journal of Geophysical Research, 111, B05401. https://doi.org/10.1029/2003JB002950

Shearer, P. M., & Orcutt, J. A. (1987). Surface and near-surface effects on seismic-waves-theory and borehole seismometer results. Bulletin of the Seismological Society of America, 77(4), 1168–1196.

Sparks, D., & Parmentier, E. (1991). Melt extraction from the mantle beneath spreading centers. Earth and Planetary Science Letters, 105(4), 368–377.

Staples, R. K., White, R. S., Brandsdottir, B., Menke, W., Maguire, P. K. H., & McBride, J. H. (1997). Faroe-Iceland ridge experiment. 1. Crustal structure of northeastern Iceland. Journal of Geophysical Research, 102(B4), 7849–7866.

Steinberger, B., & Antretter, M. (2006). Conduit diameter and buoyant rising speed of mantle plumes: Implications for the motion of hot spots and shape of plume conduits. Geochemistry, Geophysics, Geosystems, 7, Q11018. https://doi.org/10.1029/2006GC001409 Tan, Y., & Helmberger, D. V. (2007). Trans-Pacific upper mantle shear velocity structure. Journal of Geophysical Research, 112, B08301.

https://doi.org/10.1029/2006JB004853

Tharimena, S., Rychert, C., Harmon, N., & White, P. (2017). Imaging Pacific lithosphere seismic discontinuities: Insights from SS precursor modeling. Journal of Geophysical Research: Solid Earth, 122, 2131–2152. https://doi.org/10.1002/2016JB013526

Tian, Y., Shen, W. S., & Ritzwoller, M. H. (2013). Crustal and uppermost mantle shear velocity structure adjacent to the Juan de Fuca Ridge from ambient seismic noise. Geochemistry, Geophysics, Geosystems, 14, 3221–3233. https://doi.org/10.1002/ggge.20206

Tryggvason, E., & Bath, M. (1961). Upper crustal structure of Iceland. Journal of Geophysical Research, 66(6), 1913–1925. Venzke, E. e. (2013). Volcanoes of the world (Vol. 4.6.7). Washington, DC: Smithsonian Institution.

Villagomez, D. R., Toomey, D. R., Hooft, E. E. E., & Solomon, S. C. (2007). Upper mantle structure beneath the Galapagos Archipelago from surface wave tomography. Journal of Geophysical Research, 112, B07303. https://doi.org/10.1029/2006JB004672

Vinnik, L. P., Foulger, G. R., & Du, Z. (2005). Seismic boundaries in the mantle beneath Iceland: A new constraint on temperature. Geophysi-cal Journal International, 160(2), 533–538.

Weir, N. R. W., White, R. S., Brandsdottir, B., Einarsson, P., Shimamura, H., Shiobara, H., et al. (2001). Crustal structure of the northern Rey-kjanes Ridge and ReyRey-kjanes Peninsula, southwest Iceland. Journal of Geophysical Research, 106(B4), 6347–6368.

Yamamoto, M., & Morgan, J. P. (2009). North Arch volcanic fields near Hawaii are evidence favouring the restite-root hypothesis for the ori-gin of hotspot swells. Terra Nova, 21(6), 452–466.

Figure

Figure 2. Depth to the Moho from the multitaper method, map view. Bins with fewer than 3 waveforms are shaded grey
Figure 4. Waveform modeling of Sp 3-D multitaper model (black) beneath the ridge and beneath the 9 My age contour (locations of triangles in Figure 3) are compared to synthetics from forward modeling (red and green lines)
Figure 6. Hit count map for the multitaper method at 75 (left), 100 (middle), and 150 (right) km depth
Figure 8. Schematic of the evolution of lithospheric thickness beneath Iceland. The left column shows map views, 8 Ma to present at 2 Ma intervals (top to bottom) with isochons (Ivarsson, 1992)

Références

Documents relatifs

Similarly, the International Epidemiology Databases to Evaluate AIDS (IeDEA), a global consor- tium funded by the National Institutes of Health, has linked 183 clinics in 17

Identifying productive zones of the Sarvak formation by integrating outputs of different classification methods.. Pedram Masoudi, Behzad Tokhmechi, Alireza Bashari,

Phenotypic and molecular diversity of urinary isolates of Pseudomonas aeruginosa... Due to a range of mechanisms for adaptation and antibiotic resistance, PA infections are difficult

Dans une troisi`eme et derni`ere partie, nous nous int´eressons ` a des canicules intenses et ` a leurs temp´eratures extrˆemes dans le climat de la fin du 21 ` eme si`ecle en

 2007 The Authors, GJI, 169, 357–364 Journal compilation  C 2007 RAS.. Piton de la Fournaise summit area. These are the images we have used in the correlation analysis. Notice

standardized steady state heavy-duty diesel engine emission testing.. Duty Cycle for nonroad applications. This and heavier loading conditions than. )iesel Duty

A part l'Algue rouge Polysiphonia sp., chez qui la concentration en Zn est constante, et d'ail- leurs très élevée, quelle que soit la saison, dans l'ensemble,

Understanding these issues and incor- porating them in the problem definition phase is a first step toward a general filter synthesis procedure. PR